• Aucun résultat trouvé

Insights into the Headgroup and Chain Length Dependence of Surface Characteristics of Organic-Coated Sea Spray Aerosols

N/A
N/A
Protected

Academic year: 2021

Partager "Insights into the Headgroup and Chain Length Dependence of Surface Characteristics of Organic-Coated Sea Spray Aerosols"

Copied!
23
0
0

Texte intégral

(1)

HAL Id: hal-02146490

https://hal.archives-ouvertes.fr/hal-02146490

Submitted on 28 Nov 2019

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Insights into the Headgroup and Chain Length

Dependence of Surface Characteristics of

Organic-Coated Sea Spray Aerosols

S. Cheng, S. Li, N. Tsona, C. George, L. Du

To cite this version:

S. Cheng, S. Li, N. Tsona, C. George, L. Du. Insights into the Headgroup and Chain Length De-pendence of Surface Characteristics of Organic-Coated Sea Spray Aerosols. ACS Earth and Space Chemistry, ACS, 2019, 3 (4), pp.571-580. �10.1021/acsearthspacechem.8b00212�. �hal-02146490�

(2)

Insights into the head-group and chain-length dependence of surface

1

characteristics of organic-coated sea spray aerosols

2 3

Shumin Cheng,† Siyang Li,† Narcisse T. Tsona,† Christian George,‡,§ and Lin Du*,†

4 5

Environment Research Institute, Shandong University, Binhai Road 72, Qingdao 266237, China

6

School of Environmental Science and Engineering, Shandong University, Binhai Road 72, Qingdao

7

266237, China

8

§University of Lyon, Université Claude Bernard Lyon 1, CNRS, IRCELYON, F-69626 Villeurbanne,

9

France

10 11

Corresponding author: Email: lindu@sdu.edu.cn, Tel: +86-532-58631980

12 13 14

15 16

(3)

ABSTRACT

17

The structure of sea spray aerosols (SSAs) has been described as a saline core coated by organic

18

surfactants. The presence of surface-active compounds at the air-water interface can have a large

19

impact on physical, chemical and optical properties of SSAs. The surfactant molecules chosen for

20

this study, palmitic acid (PA), stearic acid (SA), arachidic acid (AA), methyl palmitate (MP), methyl

21

stearate (MS) and methyl arachidate (MA), were used to investigate the effect of alkyl chain-length,

22

head-groups and sea salts on the surface properties of these monolayers. A Langmuir trough was used

23

for measuring surface pressure−area (π−A) isotherms to reveal macroscopic phase behavior of the

24

surface films at the air-water interface. Infrared reflection absorption spectroscopy (IRRAS) was

25

employed to have a molecular-level understanding of the interfacial molecular organization. The π−A

26

isotherms indicated that sea salts, present in the subphase, exert a strong condensing effect on fatty

27

acid monolayers, while exerting expanding effect on fatty acid methyl ester monolayers, which was

28

confirmed by results from IRRAS experiments. IRRAS further revealed that the alkyl chains were in

29

an all-trans conformation, which can be evidenced by the relatively low νa(CH2) and νs(CH2)

30

stretching frequencies. The conformational order changes in the alkyl chains of different film-forming

31

species (C16 < C18 < C20) were directly revealed by analyzing the relative intensity of the νa(CH2)

32

and νs(CH2) peaks in the C-H stretching region.Thus, all the three factors alter the phase behavior

33

and molecular packing of the monolayers at the air-aqueous interface.

34 35

Keywords: Langmuir films; fatty acid; fatty acid methyl ester; sea spray aerosols; sea surface

36

microlayer

37 38

(4)

1. INTRODUCTION

39

Based on mass concentration, sea spray aerosol (SSA) is one of the largest sources of primary

40

atmospheric aerosol particles.1-2 SSAs formed at the sea surface microlayer through bubble-mediated

41

processes, are commonly composed of a sea salt core coated by a thin organic layer.3-6 This organic

42

layer is enriched with various organic species from both biological and anthropogenic sources, among

43

which, surface-active species account for a significant portion.5,7 Moreover, surface-active species in

44

the sea surface microlayer are expected to be more efficiently transferred into SSA, thus exhibiting

45

much higher concentration than that measured in the sea surface microlayer.7-9 Organic films present

46

in SSAs are reported to have inverse micelle structures with the hydrophilic head-groups toward the

47

aqueous phase and the hydrophobic tails to the air.10-11 The morphology and conformation of organic

48

films will affect aerosol growth and volatility,12 radiative absorption and scattering,6,13 reactivity with

49

atmospheric gases,10 and cloud condensation nuclei activity.3,8,14-15 Thus, understanding the surface

50

characteristics of surfactant molecules is important in atmospheric chemistry because of their

51

significant influence in physical, chemical and optical properties of SSAs.10,16

52

Previous measurements suggested that fatty acids make up a large fraction of the organic

53

materials residing at the interface of marine aerosols.4,7,17 Fatty acids present in sea surface microlayer

54

are released primarily during the lysis of phospholipid cellular membranes of marine organisms.11,14

55

Saturated fatty acids, particularly palmitic acid (PA) and stearic acid (SA), contribute significantly to

56

the organic coating of sea salt particles,11,18-19 with arachidic acid (AA) showing a relatively lower

57

abundance.20 On the other hand, a former analysis of marine aerosols collected over the

58

Mediterranean Sea using a five-stage cascade impactor, found the existence of fatty acid methyl esters

59

from C14 to C34 in SSAs, among which, methyl palmitate (MP) and methyl stearate (MS) were found

60

to be predominant.21-22 Natural sources of these fatty acid methyl esters from hydrolysis of various

61

biosynthesized esters such as triglycerides, glycolipids, phospholipids, or waxes, have been

62

reported.21 Degradation of particulate material suspended in seawater from marine organisms has

63

been found to be the source of saturated and unsaturated fatty acid methyl esters.23

64

Investigation of the surface properties of these organics at the air-water interface can provide a

65

better understanding of the chemical and physical processes taking place at the surfaces of SSAs in

66

the atmosphere. The film-forming species selected for this research are long chain fatty acids and

67

fatty acid methyl esters with alkyl chain-lengths of C16, C18 and C20. Langmuir trough has been

68

extensively used in atmospheric chemistry to understand the properties of surfaces and surfactant

69

molecules at the interface of aqueous aerosols.10,16 Surface pressure−area (π−A) isotherms of

(5)

Langmuir monolayers on aqueous surfaces are capable of revealing the underlying phase information

71

of the monolayers being subject to constant compression. However, to gain molecular-level insights

72

into the monolayers, spectroscopic techniques are needed. Infrared reflection absorption spectroscopy

73

(IRRAS) is a helpful technique in studying surface phenomena. Possessing the advantage of being a

74

fast and nondestructive technique,24 IRRAS is sensitive to Langmuir monolayers if sufficient scans

75

are taken to obtain spectra with good signal-to-noise ratios.25 The application of IRRAS enables us

76

to acquire information about molecular structure, conformation, and orientation of the film-forming

77

species.25-27 There are extensive studies elucidating the microscopic profile of Langmuir films of

78

saturated fatty acids at the air-water interface using IRRAS.4,28-30 In addition, fatty acid methyl esters

79

are also known to be able to form organized structures at the air-water interface.31-34 Previous studies

80

of saturated long chain fatty acid and fatty acid methyl ester monolayers were carried out at the

air-81

water interface with IRRAS technique,by which the conformational order of the monolayers were

82

shown to be increased with increasing alkyl chain-length.35-36 The response of SA and AA Langmuir

83

monolayers to atmospheric inorganic ions was explored by Langmuir and IRRAS methods, which

84

confirmed the existence of inorganic ions in the fatty acid monolayers and its impact on the surface

85

properties of aqueous-phase aerosols.15 Former investigations were mainly conducted on pure water

86

(PW) or ion-containing subphases. However, properties of sea surface relevant surfactant monolayers

87

on the artificial seawater (ASW) subphase have been rarely studied.37 The ASW used in this work is

88

pure water enriched with relevant sea salts, which is more representative of the environment that the

89

surfactant monolayers are exposed to.

90

In the present study, fatty acid/fatty acid methyl ester-ASW systems were chosen as proxies to

91

further understand the surface properties of marine aerosols. To investigate the influence of sea salts

92

on the studied monolayers, PW was also used as aqueous phase for comparison. Phase behavior and

93

molecular-level features of the fatty acid (PA, SA, AA) and fatty acid methyl ester (MP, MS, MA)

94

monolayers at the air-aqueous interface were examined using π−A isotherms and IRRAS spectra.

95

These compounds commonly have a carboxylic or methyl ester head-group connected to a saturated

96

hydrocarbon chain with different chain-lengths. The objective of the experiments outlined in this

97

paper is to determine the effect of alkyl chain-length, head-groups and sea salts on the properties of

98

the surfactant monolayers at the air-aqueous interface.

99 100

(6)

2. EXPERIMENTAL SECTION

101

2.1 Materials. PA (≥98%, Adamas-beta), SA (98%, Aladdin), AA (99%, Aladdin), MP (99%,

102

Aladdin), MS (99%, Aladdin) and MA (≥98%, Aladdin) were used without further purification. These

103

chemicals were dissolved in chloroform to a final concentration of 1 mM. Ultrapure water with a

104

resistivity of 18.2 MΩ was obtained from a Millipore Milli-Q purification system. ASW (see Table

105

1 for the detailed composition and concentrations) is a ten components mixture with a total

106

concentration of approximately 0.53 M.37-38 Specifically, it consists of: NaCl (≥99%, Acros Organics),

107

Na2SO4 (99%, Alfa Aesar), KCl (3 M, Alfa Aesar), NaHCO3 (≥99.7, Alfa Aesar), KBr (≥99%, Alfa

108

Aesar), H3BO3 (99.5%, Innochem), NaF (≥99%, Acros Organics), MgCl2·6H2O (≥99, Aladdin),

109

CaCl2·2H2O (99%, Adamas-beta), SrCl2·6H2O (≥99, Alfa Aesar). All these salts were used as

110

received. The pH of the ASW was measured in the range of 8.0 ± 0.2, a value representative of the

111

real seawater. The prepared ASW solution was allowed to equilibrate for several hours before

112

experiments were conducted.

113 114

Table 1. Composition of the artificial seawater

115

Type of cation Salt Concentration in aqueous solution (mM) monovalent NaCl 426 Na2SO4 29.4 KCl 9.45 NaHCO3 2.43 KBr 0.857 H3BO3 0.438 NaF 0.0744 divalent MgCl2·6H2O 55.5 CaCl2·2H2O 10.8 SrCl2·6H2O 0.0937 116

2.2 Monolayer Spreading and Isotherm Measurements.

117

The surface pressure π is a measurement of the difference between the surface tension when the

118

surface is covered in a surfactant and the surface tension of the bare surface (π = γ0 - γLangmuir, where

119

γ0 is the surface tension of pure water and γLangmuir is the surface tension of water with the Langmuir

120

film at the air-water interface).16,39 The surface tension value of ASW was calculated to be in the

121

range of 73.7-74.0 mN/m at 291 ± 1 K (details are given in the Supporting Information). Standard

122

deviations of the molecular area and surface pressure were ±1 Å2/molecule and ±0.5 mN/m,

123

respectively.

(7)

π−A isotherms of the monolayers at the air-aqueous interface were recorded using a

computer-125

controlled Langmuir trough with two movable barriers sitting on top of the aqueous surface. The

126

trough is made out of Teflon and has inside dimensions of 65 mm × 280 mm × 3 mm. It was placed

127

on a vibration isolation table and closed in aPlexiglas box. At the beginning of the experiments, the

128

two barriers were placed at the ends of the trough.Tens of microliters of chloroform solutions of fatty

129

acids or fatty acid methyl esters were spread dropwise onto PW or ASW subphase using a glass

130

microsyringe. After deposition, about 15 min was allowed before compression to permit the solvent

131

to evaporate and the film to spread spontaneously. A pressure sensor with a Wilhelmy plate made

132

from a piece of rectangular filter paper was used to monitor the surface pressure with high sensitivity.

133

The π−A isotherms were obtained with the pressure sensor while the surface area available for the

134

surfactant molecules was decreased between the barriers. The monolayer at the air-aqueous interface

135

was continuously compressed at a constant rate of 3 mm/min. All experiments were performed at

136

ambient temperature (291 ± 1 K). Each experiment was run at least three times to ensure

137

reproducibility.

138 139

2.3 IRRAS Measurements. IRRAS spectra of the monolayers were recorded on a Bruker Vertex 70

140

FTIR spectrometer equipped with an external variable angle reflectance accessory for monolayer

141

measurements. To have maximum signal strength, the incidence angle of the IR beam was set at 40°

142

with respect to the surface normal. IRRAS spectra were collected over the range of 4000-400 cm-1 by

143

using a liquid-nitrogen cooled HgCdTe (MCT) detector and averaged for 2000 scans at a resolution

144

of 8 cm-1. For IRRAS spectra collection, the monolayers were continuously compressed to a desired

145

surface pressure from ∼0 mN/m. When the barriers were stopped, IRRAS spectra were obtained after

146

a time delay of 60 s, allowed for film equilibrium between trough movement and data collection.

147

IRRAS spectra were obtained at the surface pressure of 28 mN/m, which corresponded to the untilted

148

condensed phase of the π−A isotherms. During the IRRAS data collection, surface pressure changed

149

slightly for the monolayers (≤0.2 mN/m).

150 151

3. RESULTS AND DISCUSSION

152

3.1 Surface Pressure−Area Isotherm. π−A isotherms provide information about the phase behavior

153

of the monolayers at air-water interfaces. During the movement of the barriers, the phase of the

154

surface monolayer changes with the increase of the surface pressure. The phase changes can thus be

155

recognized from characteristics of the π−A isotherms. The surface pressure monitored as a function

(8)

of surface area of fatty acid and fatty acid methyl ester monolayers at room temperature are shown in

157

Figure 1. These π−A isotherms show remarkable changes along with alkyl chain-length, head-groups,

158

and aqueous subphases.

159 160

161

Figure 1. Surface pressure−area isotherms of fatty acid ((a), (b)) and fatty acid methyl ester ((c), (d))

162

monolayers on pure water (PW) and artificial sea water (ASW) subphases.

163 164

It can be seen from Figure 1(a) that the monolayers of fatty acids on pure water (PW) subphase

165

exhibit the following typical features in π−A isotherms upon compression. At low surface pressure

166

(π = 0 mN/m), the fatty acid monolayers show a gaseous-tilted condensed (G-TC) coexistence phase

167

before the lift-off area. In this phase, the alkyl chains are mostly free in space.40 And then, the

168

monolayers were enforced into a tilted condensed (TC) phase after subsequent compression. In this

169

phase, there are less spatial movements for the fatty acid molecules. Further compression results in a

170

kink indicates a second-order phase transition from TC to an untilted condensed (UC) phase. In the

171

UC phase the hydrocarbon chains of fatty acids are almost perpendicular to the water subphase.16 In

(9)

addition, the surface pressure of second-order phase transitions from TC to UC phase upon

173

compression occur at about 24 mN/m, which is consistent with previous studies.25,41-42 Finally,

174

compressing the monolayer even further leads to a collapse state where the monolayer forms

three-175

dimensional structures because the surface is no longer stable.16,43-44

176

The π−A isotherms of fatty acids obtained on ASW subphase are presented in Figure 1(b). The

177

isotherm conducted solely with the ASW subphase is shown in Figure S1, from which can be seen

178

that the surface pressure fluctuates around 0 mN/m, indicating that no or negligible surface active

179

substance is present in the ASW subphase. Thus, the influence of organic contaminants present in sea

180

salts can be ruled out. In the presence of sea salts, some changes are found in comparison with the

181

isotherms obtained on the PW subphase. It can be clearly seen that the TC phase disappears, as well

182

as the second-order phase transition. Beyond the lift-off points, the surface pressure increases steeply

183

with a formation of UC phase until the collapse of the monolayers occurs. This is in agreement with

184

a previous report studying the π−A isotherms for PA monolayers formed on CaCl2 solution in.25 This

185

behavior was attributed to the condensing effect of metal cations as a result of forming fatty acid

186

salts.25,45 Thus, a more orderly packed structure can be speculated for fatty acid monolayers on ASW

187

subphase. The differences in properties of π−A isotherms between Figures 1(a) and 1(b) indicate that

188

sea salts present in the subphase alter the macroscopic phase behavior of the long chain fatty acid

189

monolayers.

190

A strong chain-length dependence of fatty acids is also observed, as the lift-off areas are

191

concerned, which becomes gradually smaller with increasing chain-length. The lift-off areas of the

192

isotherms on PW are 30.3, 29.3 and 24.2 Å2/molecule for PA, SA and AA, respectively. The sequence

193

and general shape of the interfacial isotherms herein are consistent with previous reports.37,46-47 The

194

chain-length dependence of lift-off areas indicates that the van der Waals energy increases with the

195

length of chain, and draws the molecules closer as the intermolecular attraction increases.47-48 On the

196

other hand, compared to PW, condensation of the fatty acid monolayers occurs when ASW is used

197

as subphase as illustrated in Figure 1(b). It is evident that with ASW as subphase, the lift-off areas

198

are decreased for individual fatty acids. The change in lift-off areas for PA, SA and AA are -1.4, -2.2,

199

-1.7 Å2/molecule, respectively. Consequently, all the studied fatty acid molecules become more

200

densely packed on ASW than on PW.46 At the air-ASW interface, the favorable electrostatic

201

interaction or complexation between fatty acids and sea salts leads to a denser chain packing relative

202

to that of air-PW interface.49

203

Being molecules with saturated single alkyl chains like fatty acids, fatty acid methyl ester

(10)

molecules are fairly compressible as well, and ultimately, can be packed in a highly ordered structure

205

at high surface pressures. As can be seen from Figures 1(c) and 1(d), the slight TC-UC transition

206

occurs at about 8 mN/m as the surface pressure rises for all the three fatty acid methyl ester

207

monolayers, irrespective of the subphase. Specifically, the π−A isotherm obtained for MP in this

208

study is consistent with a previous research about the temperature dependence of MP monolayers,

209

which found that they are fully condensed below 293 K and no plateau can be observed in the

210

isotherm.35,50 Unlike fatty acid monolayers, the presence of sea salts does not change significantly the

211

shape of fatty acid methyl ester isotherms, although a clear shift to larger mean molecular areas can

212

be observed. Thus, sea salts exert an opposite effect on fatty acid methyl esters compared to fatty

213

acids, resulting in an expansion of the fatty acid methyl ester monolayers. The lift-off areas are clearly

214

observed at about 34.6 and 25.4 and 24.4 Å2/molecule for MP, MS and MA monolayers on PW,

215

respectively. However, when ASW is used as subphase, the values shift to about 45.5, 29.3 and 26.3

216

Å2/molecule. Similarly, a former investigation of MP on PW and NaCl solutions with different ion

217

concentrations (1, 2 and 3 M), found the expanding effect of NaCl on the MP monolayer from the

218

increasing lift-off area with increasing NaCl concentration.50 The reason of this behavior was

219

explained to be the increased ionic strength, which influences interactions of the water molecules

220

with the carboxyl group.

221

To sum up, the presence of sea salts in the subphase has opposite effects on the fatty acid and

222

fatty acid methyl ester monolayers. In addition, similar with fatty acids, the π−A isotherms of fatty

223

acid methyl esters shift to smaller molecular areas with increasing chain-length. Hence, further

224

increase in the hydrocarbon chain-length results in more densely packing of the fatty acid methyl

225

ester molecules. Thus, the conformational order of the alkyl chains increases with increasing

chain-226

length can be concluded from the π−A isotherms: C16 < C18 < C20, irrespective of head-groups or

227

subphases.

228

When comparing Figures 1(b) and 1(d), the difference in π−A isotherms on ASW subphase

229

indicates that the phase sequence is changed because of the esterification of the carboxyl group, with

230

the appearance of TC phase and second-order phase transitions in fatty acid methyl ester monolayers.

231

When ASW is used as subphase, it can be seen from Figures 1(b) and 1(d) that the lift-off areas of

232

fatty acid methyl esters are much larger than corresponding fatty acids. The main difference between

233

the fatty acids and the fatty acid methyl esters is the head-group structure, where the acids are

234

sensitively influenced by the subphase pH. Empirical evidence suggests that under the conditions of

235

about pH=8.2, the carboxyl group is partially dissociated to be negatively charged.4,25,51 It is likely

(11)

that the stability and surface activity of these long chain fatty acids decrease upon dissociation of the

237

carboxylic acid proton.4 However, the stability of these monolayers can be greatly improved by

238

electrostatic attractions or complexation with various sea salts in the aqueous subphase.49,52 By means

239

of IRRAS, the surface propensity of PA molecules was found to be increased by adding NaCl into

240

the subphase,4 suggesting that deprotonated fatty acids may be found at the air-aqueous interface due

241

to the role of sea salts in surface stabilization. A study of SA monolayers was carried out on 1, 10 and

242

100 times diluted ASW.37 The π−A isotherms of the SA monolayers show an enhanced stability of

243

the film against fracture when the sea salt concentration of the subphase was higher. In case of fatty

244

acid methyl esters, no effect is expected due to ionization of head-groups, because the pH of the

245

subphase is too low to observe any measurable hydrolysis of the esters. The observed different trends

246

in the isotherms may in part be due to the different interaction mechanisms between the sea salts and

247

the head-groups.

248

It has been well documented that at low surface pressure, the methyl ester head-group is

E-249

configured for expanded fatty acid methyl ester monolayers, where substantial part of the oxo-methyl

250

group is pointing out of the water (Figure S2). The E isomer of fatty acid methyl esters allows the

251

hydration of the polar group and hinders the electrostatic interactions between hydrophilic head-group

252

and cation. However, at higher surface pressure, the head-group is Z-configured with the oxo-methyl

253

component directed into the water for more orderly packed compressed states.35,53 In addition, several

254

studies have described the expulsion of water molecules from the monolayer as surface pressure

255

increases.32-33,35 When the fatty acid methyl ester monolayers are in Z-configuration, the carbonyl

256

group is shielded by the oxo-methyl component and, thus, the carbonyl group is more or less

257

prevented from being hydrogen-bridged by water molecules. The expelled water from fatty acid

258

methyl ester films could affect the arrangement of Z isomer, allowing the cations to penetrate into the

259

film (Figure S3). In this way, the growth of three-dimensional structures could be favored because

260

the partial charge of the head-group of fatty acid methyl ester would be compensated.33 Therefore,

261

we can speculate that the main cause of the instability of the fatty acid methyl ester monolayers in the

262

presence of sea salts is a collapse process which involves the formation of three-dimensional nuclei

263

on the monolayer surface.

264

The above descriptions suggest that both interactions between adjacent amphiphiles and

265

interactions between amphiphiles and the subphase are important to the macroscopic phase behavior

266

of the monolayers. However, to get deep insights into microscopic molecular arrangement and

267

underlying mechanisms, advanced spectroscopic techniques are necessary. This consideration led us

(12)

to explore the molecular conformation of the monolayer films by means of the IRRAS technique.

269 270

3.2 IRRAS Spectra. The monolayers of fatty acids with different chain-lengths (PA, SA, and AA)

271

and corresponding fatty acid methyl esters (MP, MS, MA) on air-aqueous interface were examined

272

using IRRAS. This technique enables us to probe microscopic information such as the conformation

273

order and orientation of the monolayers at the molecular-level. The vibrational modes investigated

274

encompass both head and tail groups of the sample molecules. The stretching vibrations of C-H

275

(ν(CH2)) at 2820-2950 cm-1, the ν(COO) and scissoring band of C-H (δ(CH2)) at 1400-1500 cm-1,

276

and the ν(C=O) at 1700-1800 cm-1 were probed.

277 278

279

Figure 2. IRRAS spectra (2820-2950 cm-1) of the fatty acid ((a), (b)) and fatty acid methyl ester ((c),

280

(d))monolayers recorded at 28 mN/m on pure water (PW) and artificial sea water (ASW) subphases

281

at the incidence angle of 40°.

282 283

IRRAS spectra (2820-2950 cm-1) of the fatty acid and fatty acid methyl ester monolayers in the

284

UC phase on PW and ASW subphases are shown in Figures 2(a)-(d). In Figure 2(a), for fatty acid

285

monolayers on air-PW interface, the two bands at around 2916 cm-1 and 2850 cm-1 can be assigned

286

to the methylene antisymmetric (νa(CH2)) and methylene symmetric (νs(CH2)) stretching vibrations

(13)

of the hydrocarbon chains, respectively. At the incidence angle of 40°, these bands show negative

288

reflection absorbance. The νa(CH2) and νs(CH2) frequencies have been known to be sensitive to the

289

conformation order of hydrocarbon chains.45,54 Lower frequencies are characteristic of preferential

290

all-trans conformers in highly ordered chains, while the number of gauche conformers increases with 291

increasing frequency and width of the bands.17,29 For all-trans conformations of the fully extended

292

tail chains, the symmetric and asymmetric stretching vibrations of the methylene groups are usually

293

present in the narrow ranges of 2846-2850 and 2915-2918 cm-1, respectively, and in the distinctly

294

different ranges of 2854-2856 and 2924-2928 cm-1 for disordered chains characterized by a significant

295

presence of gauche conformations.55 The relatively low frequency positions of the νs(CH2) and

296

νa(CH2) stretching modes at about 2850 cm-1 and 2916 cm-1, indicate that the alkyl chains are mostly

297

in highly ordered all-trans conformations.56 This shows clearly that the alkyl chains are almost

298

perpendicular to the air-water interface. The all-trans conformation can also be found in other

299

compressed monolayers shown in Figures 2(b), 2(c) and 2(d), irrespective of the head-groups or the

300

subphases. The highly ordered structure of fatty acids directly correlates with the van der Waals

301

interaction between adjacent alkyl chains, the interactions between adjacent COOH head-groups and

302

those between head-groups and aqueous subphase. An orderly packed structure can maximize

303

interactions with an all-trans conformation between adjacent alkyl chains.25 For the fatty acid methyl

304

ester monolayers, the steric demand in the E-configuration is very high. This structure results in quite

305

strongly tilted molecules and a poor conformational order. When the monolayer is further compressed,

306

a reduction of the available area forces molecules to approach each other and pack more densely. As

307

a compromise, the methyl group is squeezed into the subphase, thus resulting in a Z-conformation of

308

amphiphiles in the condensed phase. With this conformation, the oxo-methyl group facilitates the

all-309

trans configuration of the alkyl chains. Consequently, the head-group can be forced into the Z-310

configuration with increasing surface pressure and result in the all-trans conformation of the alkyl

311

chains.35

312

All the fatty acids and corresponding methyl esters commonly possess saturated hydrocarbon

313

chains with different chain-lengths. IRRAS bands arising from these alkyl chains provide the clear

314

spectra and hence reliable information about molecular conformation in the monolayers.30 The peak

315

heights and areas of the νa(CH2) and νs(CH2) bands indicate the packing density of the alkyl chains.35

316

It can be evidenced from Figure 2 that the peak heights and areas for the methylene stretching

317

vibrations increase with increasing chain-length, irrespective of head-groups or subphases. As the

318

directions of νa(CH2) and νs(CH2) vibrational modes are orthogonal to the molecular axis, strong

(14)

intensities of the bands indicate that the molecule stands nearly perpendicular to the water subphase

320

when the hydrocarbon chain is in the all-trans conformation.57 Hence, the much smaller peak

321

intensity of C16 relative to the higher homologues is indicative of a substantially stronger tilt and less

322

ordered conformation of the molecules, reflecting more orderly packed structure of higher

323

homologues at the UC state. It is necessary to consider these data in relation to the π−A isotherms of

324

Figure 1, which shows that smaller areas were occupied by monolayers formed by surfactants with

325

longer alkyl chain-length.

326

Table 2. The peak-height intensity ratio between the antisymmetric and symmetric bands of the CH2

327

groups (Ias/Is) for the fatty acid and fatty acid methyl ester monolayers on pure water (PW) and

328

artificial sea water (ASW) subphases.

329 330

331

The conformational order changes in the alkyl chains introduced by sea salts, head-groups or

332

alkyl chain-length can be further revealed by analyzing the relative intensity of the νa(CH2) and the

333

νs(CH2) peaks in the C-H stretching region.15 The peak-height intensity ratios between the

334

antisymmetric and symmetric bands of the CH2 groups (Ias/Is) for the studied monolayers are

335

presented in Table 2 for direct comparison. In Figure 2, the intensities of νs(CH2) peaks are relatively

336

weaker than those of the νa(CH2) peaks, thus giving Ias/Is values greater than one. Qualitatively

337

speaking, larger Ias/Is ratio indicates more orderly packed alkyl chains with nearly all-trans

338

conformation.58-60 It can be seen from Table 2 that the ratio values are smaller for monolayers formed

339

by molecules with shorter alkyl chain-length, irrespective of head-groups or subphases, indicating the

340

existence of gauche defects in corresponding monolayers. Therefore, the pronounced chain order

341

increase with increasing chain-length can be concluded from IRRAS spectra: C16 < C18 < C20,

342

which is in line with the conclusion obtained from the π−A isotherms (Figure 1). As was shown in

343

the π−A isotherms, sea salts demonstrate a condensing effect on the fatty acid monolayers, which

344

consequently leads to the absence of the TC phase. The IRRAS spectra obtained on the ASW surface

345

confirm this effect, as can be seen from the larger intensity ratios of the νa(CH2) over the νs(CH2) than

346

those on the PW subphase in individual spectra, with values of Ias/Is increasing from 1.10, 1.23 and

347

1.43 to 1.19, 1.32 and 1.78, respectively. This result can be attributed to the decrease in the

348

subphase PA SA AA MP MS MA

PW 1.10 1.23 1.43 1.21 1.29 1.33 ASW 1.19 1.32 1.78 1.14 1.19 1.26

(15)

concentration of gauche defects and tilt angle of monolayers formed on ASW subphase. With respect

349

to fatty acid methyl esters, the intensity ratios of Ias/Is obtained on ASW (1.14, 1.19 and 1.26) surface

350

are smaller than those on PW (1.21, 1.29 and 1.33), which indicates that the fatty acid methyl ester

351

monolayers are disordered by sea salts. This can be evidenced by the expanding effect introduced by

352

sea salts on fatty acid methyl ester monolayers. Therefore, IRRAS experiments confirm the contrary

353

effects of sea salts on fatty acid and fatty acid methyl ester monolayers as can be observed from π−A

354

isotherms.

355

Evidences of monolayer orientation and structural changes along with alkyl chain-length,

head-356

groups, and subphases are provided mainly by details of the νa(CH2) and the νs(CH2) bands. Peak

357

position, height, area and intensity ratios were utilized to support the analysis. These characteristics

358

correlate well with those shown by the π−A isotherms. Information about the dependence of the chain

359

order on the alkyl chain-length and subphases, i.e., the overall effect of an increase in alkyl

chain-360

length leading to an increase in order, ASW acts to condense fatty acid films and expand fatty acid

361

methyl ester films, was inferred from both the IRRAS spectra and the π−A isotherms. The IRRAS

362

technique not only allows for the characterization of all the above chain conformation and orientation

363

details, but also provides valuable information about molecular interaction between the monolayers

364

and the aqueous subphase.

365 366

(16)

Figure 3. IRRAS spectra (1380-1800 cm-1) of the fatty acid ((a), (b)) and fatty acid methyl ester ((c),

368

(d)) monolayers recorded at 28 mN/m on pure water (PW) and artificial sea water (ASW) at the

369

incidence angle of 40°.

370 371

Figure 3 shows IRRAS spectra (1800-1380 cm-1) of the ν(C=O), δ(CH2) and ν(COO) regions of

372

the fatty acid and fatty acid methyl ester monolayers in the UC phase on the PW and ASW subphases.

373

Three peaks are observed at 1739, 1720, and 1704 cm-1, which can be attributed to the stretching

374

vibrations of the free C=O group, the C=O group involved in one and two hydrogen bonds,

375

respectively. A previous IRRAS investigation of SA monolayer at the air-water interface gave similar

376

results.36 The ability of the carbonyl group to form hydrogen bonds was explained to be a mixture

377

effect of hydration by the water subphase and by side-bridging hydrogen-bond formation between

378

adjacent fatty acid molecules.36,61 The sharp singlet observed at 1472 cm-1 is ascribed to the δ(CH2)

379

band of the methylene groups.29,54

380

IRRAS spectra of fatty acids on the ASW subphase (Figure 3 (b)) in the region of COO

381

stretching vibrations can provide insights into interaction mechanisms between carboxylic acid

head-382

groups and the subphase. In the presence of sea salts, additional peaks arising from the asymmetric

383

(νa(COO) and symmetric (νs(COO)) stretching modes of the COO group are observed. The three

384

peaks including 1558, 1542, 1523 cm-1 are resulted from the splitting of the νa(COO) stretching

385

vibration,15 while the peak at 1419 cm-1 is assigned to the νs(COO) stretching mode.38 Complexation

386

of ions to surface-active species has been known to alter their orientation, packing, and surface

387

morphology.17,62 The stability of the SA monolayer was found to be increased significantly at high

388

pH values due to ionization of the surfactant by Ca2+ and Mg2+ in the subphase.63 It has been reported

389

that metal cations can bind to the carboxylate group in several ways including ionic binding,

390

unidentate type, bidentate chelate type and bidentate bridging type.24 The bonding type of metal

391

cations to the carboxylate group in the UC state can be classified by the difference between the

392

antisymmetric and symmetric COO stretching frequencies. In this work, the differences in νa(COO)

393

and νs(COO) stretching frequencies give three values of about 139, 123 and 104 cm-1, respectively.

394

The difference in νa(COO) and νs(COO) stretching frequencies for dissociated acid was estimated to

395

be 138 cm-1.24 Typically, the values for bidentate bridging coordination are somewhat close to that

396

for a dissociated carboxylate ion, and values of bidentate chelate coordination are less than that of a

397

dissociated carboxylate ion.24 Thus, the main component at 1558 cm-1 belongs to a bidentate bridging

(17)

structure, while the components at 1542 cm-1 and 1523 cm-1 can be attributed to bidentate chelate

399

coordinations. Hence, in the presence of the ASW subphase, dissociated fatty acids form bidentate

400

bridged and bidentate chelate coordinations. As the ASW used herein is a complex mixture of sea

401

salts, it is hard to distinguish which cation the fatty acids are binding to. In this regard, sea salts are

402

treated as a whole to consider their interaction with the head-groups of fatty acid and fatty acid methyl

403

ester molecules.

404 405

3.3 Atmospheric Implications. The organic films that reside at the air-water interface exert a

406

significant impact on many properties of SSAs, such as its ability to exchange species including water

407

molecules and traces gases across the interface,64 its ability to absorb or scatter radiation,6,10,13 and its

408

reactivity towards oxidative gases.65 Long chain fatty acid and fatty acid methyl ester monolayers at

409

the air-aqueous interface were utilized as simplified model of organic-coated SSAs. The impact of

410

sea salts, head-groups and alkyl chain-length on phase behavior and molecular organization of the

411

monolayer films was fully characterized. The higher stability of monolayers formed by species with

412

longer alkyl chain-length is clear. The lifetime of the hydrophobic layer on SSA is dependent on many

413

variables, an important one being the carbon chain-length of the surfactants comprising the coating.

65-414

66 Properties of the aqueous core, including pH and composition, also affect the stability of the organic

415

surface films. Adding sea salts into the subphase improves the stability of the fatty acid films by

416

binding to the carboxylic acid groups through bidentate bridged and bidentate chelate coordinations.

417

Thus, deprotonated fatty acids may be found at the air-aqueous interface of aerosol particles partly

418

due to the role of sea salts in surface stabilization.

419

ASW caused condensation of the fatty acid surface films, leading to tightly packed molecules.

420

However, the expansion effect was introduced by ASW toward fatty acid methyl ester films, which

421

led to loosely packed molecules. Thus, we can speculate that fatty acid molecules reside at the

422

interface of SSAs with greater stability and higher packing density relative to fatty acid methyl esters.

423

This is in line with field measurements of marine aerosols utilizing time-of-flight secondary ion mass

424

spectrometry (TOF-SIMS) as a surface sensitive analysis technique. The aerosols collected in the

425

field exhibited surface layers dominated by fatty acids.67 One major effect of surface active organic

426

monolayer shown both by observations and modeling, is the lowering of the particle surface

427

tension.39,68 Surface active species present at the air-water interface have the potential to lower the

428

surface tension of a growing droplet relative to pure water at a given relative humidity.69 A lower

429

surface tension promotes small particle growth at lower relative humidity in accord with the Kelvin

(18)

effect69-70 and increasing particle cloud condensation nuclei activation efficiency,39,68 thereby causing

431

the droplets to grow larger than predicted. As can be seen from the surface pressure-area isotherms,

432

beyond the lift-off points, the surface pressure of the interface increases with decreasing mean

433

molecular areas, indicating the reduction of surface tension. The surface tension of ASW interface

434

covered by fatty acid reduces more rapidly along with deceasing mean molecular areas than in

435

corresponding methyl ester. Thus the effectiveness of the film on surface tension reduction will

436

depend on the species of film-forming molecules as well as chemical composition of the aqueous

437

core. In addition, the optical properties of aerosols are largely dependent upon their size and in this

438

regard, the alteration in aerosol size will affect their scattering efficiency.46

439 440

4. CONCLUSIONS

441

In this work, monolayers of long chain fatty acids and fatty acid methyl esters (C16, C18, C20)

442

at the air-aqueous interface were used as proxies for the organic-coated SSAs. Both π−A isotherms

443

and IRRAS spectra were applied to systematically investigate the effect of alkyl chain-length,

head-444

groups and sea salts on the surface properties of organic monolayers. It was shown by π−A isotherms

445

that sea salts have a condensing effect on fatty acid monolayers, meanwhile, obvious differences in

446

phase behavior were detected over the PW and ASW subphases. However, an expansion effect of sea

447

salts on fatty acid methyl ester monolayers was observed, without any distinct change of the phase

448

transitions between π−A isotherms detected over the PW and ASW subphases. The pronounced chain

449

order increase with increasing chain-length (C16 < C18 < C20) was revealed by π−A isotherms,

450

irrespective of head-groups or subphases. These findings were further confirmed by IRRAS spectra.

451

Substantial intensity ratio increases in the Ias/Is were observed on monolayers formed by species with

452

longer chain-length. From the differences between νa(COO) and νs(COO) stretching frequencies, the

453

dominant binding coordinations between deprotonated fatty acids and sea salts were found to be

454

bidentate bridging and bidentate chelate. These results indicate that the surface characteristics of

455

organic-coated SSAs are influenced by both the chemical composition of the aqueous core and

456

species of film-forming molecules.

457 458 ASSOCIATED CONTENT 459 Supporting Information 460

Calculation of surface tension of artificial seawater, surface pressure−area isotherms of artificial

461

seawater and isotherm of steric acid monolayer on artificial seawater subphase (Figure S1), schematic

(19)

representation of E- and Z-conformation of fatty acid methyl esters at the air-water interface (Figure

463

S2), and schematic representation of fatty acid methyl ester monolayers at the air-seawater interface

464 (Figure S3). 465 466 AUTHOR INFORMATION 467 Corresponding Author 468

*Email: lindu@sdu.edu.cn, Tel: +86-532-58631980

469

Notes

470

There are no conflicts of interest to declare.

471 472

ACKNOWLEDGMENTS

473

This work was supported by National Natural Science Foundation of China (91644214, 21876098),

474

Shandong Natural Science Fund for Distinguished Young Scholars (JQ201705) and the the Marie

475

Curie International Research Staff Exchange project MARSU (Grant 690958).

476 477

REFERENCE

478

(1) Braun, R. A.; Dadashazar, H.; MacDonald, A. B.; Aldhaif, A. M.; Maudlin, L. C.; Crosbie, E.;

479

Aghdam, M. A.; Mardi, A. H.; Sorooshian, A. Impact of wildfire emissions on chloride and bromide

480

depletion in marine aerosol particles. Environ. Sci. Technol. 2017, 51 (16), 9013-9021.

481

(2) Quinn, P. K.; Coffman, D. J.; Johnson, J. E.; Upchurch, L. M.; Bates, T. S. Small fraction of

482

marine cloud condensation nuclei made up of sea spray aerosol. Nat. Geosci. 2017, 10 (9), 674-679.

483

(3) Jayarathne, T.; Sultana, C. M.; Lee, C.; Malfatti, F.; Cox, J. L.; Pendergraft, M. A.; Moore, K. A.;

484

Azam, F.; Tivanski, A. V.; Cappa, C. D.; Bertram, T. H.; Grassian, V. H.; Prather, K. A.; Stone, E.

485

A. Enrichment of saccharides and divalent cations in sea spray aerosol during two phytoplankton

486

blooms. Environ. Sci. Technol. 2016, 50 (21), 11511-11520.

487

(4) Adams, E. M.; Wellen, B. A.; Thiraux, R.; Reddy, S. K.; Vidalis, A. S.; Paesani, F.; Allen, H. C.

488

Sodium-carboxylate contact ion pair formation induces stabilization of palmitic acid monolayers at

489

high pH. Phys. Chem. Chem. Phys. 2017, 19 (16), 10481-10490.

490

(5) Tseng, R. S.; Viechnicki, J. T.; Skop, R. A.; Brown, J. W. Sea-to-air transfer of surface-active

491

organic-compounds by bursting bubbles. J. Geophys. Res.: Oceans 1992, 97 (C4), 5201-5206.

492

(6) Tervahattu, H.; Hartonen, K.; Kerminen, V. M.; Kupiainen, K.; Aamio, P.; Koskentalo, T.; Tuck,

493

A. F.; Vaida, V. New evidence of an organic layer on marine aerosols. J. Geophys. Res. 2002, 107

494

(D7-D8), AAC1-AAC9.

495

(7) Cochran, R. E.; Laskina, O.; Jayarathne, T.; Laskin, A.; Laskin, J.; Lin, P.; Sultana, C.; Lee, C.;

496

Moore, K. A.; Cappa, C. D. Analysis of organic anionic surfactants in fine and coarse fractions of

497

freshly emitted sea spray aerosol. Environ. Sci. Technol. 2016, 50 (5), 2477-2486.

498

(8) Lin, W.; Clark, A. J.; Paesani, F. Effects of surface pressure on the properties of Langmuir

499

monolayers and interfacial water at the air-water interface. Langmuir 2015, 31 (7), 2147-2156.

500

(9) Tervahattu, H.; Hartonen, K.; Kerminen, V. M.; Kupiainen, K.; Aarnio, P.; Koskentalo, T.; Tuck,

501

A. F.; Vaida, V. New evidence of an organic layer on marine aerosols. J. Geophys. Res.: Atmos. 2002,

502

107 (D7), 4053. 503

(20)

(10) Donaldson, D. J.; Vaida, V. The influence of organic films at the air-aqueous boundary on

504

atmospheric processes. Chem. Rev. 2006, 106 (4), 1445-1461.

505

(11) Ellison, G. B.; Tuck, A. F.; Vaida, V. Atmospheric processing of organic aerosols. J. Geophys.

506

Res.: Atmos. 1999, 104 (D9), 11633-11641. 507

(12) Feingold, G.; Chuang, P. Y. Analysis of the influence of film-forming compounds on droplet

508

growth: Implications for cloud microphysical processes and climate. J. Atmos. Sci. 2002, 59 (12),

509

2006-2018.

510

(13) Vaida, V. Atmospheric radical chemistry revisited Sunlight may directly drive previously

511

unknown organic reactions at environmental surfaces. Science 2016, 353 (6300), 650-650.

512

(14) Zhang, T.; Cathcart, M. G.; Vidalis, A. S.; Allen, H. C. Cation effects on phosphatidic acid

513

monolayers at various pH conditions. Chem. Phys. Lipids 2016, 200, 24-31.

514

(15) Li, S. Y.; Du, L.; Wei, Z. M.; Wang, W. X. Aqueous-phase aerosols on the air-water interface:

515

Response of fatty acid Langmuir monolayers to atmospheric inorganic ions. Sci. Total Environ. 2017,

516

580, 1155-1161. 517

(16) Larsen, M. C. Binary phase diagrams at the air-water interface: An experiment for undergraduate

518

physical chemistry students. J. Chem. Educ. 2014, 91 (4), 597-601.

519

(17) Adams, E. M.; Casper, C. B.; Allen, H. C. Effect of cation enrichment on

520

dipalmitoylphosphatidylcholine (DPPC) monolayers at the air-water interface. J. Colloid Interface

521

Sci. 2016, 478, 353-364. 522

(18) Rouviere, A.; Ammann, M. The effect of fatty acid surfactants on the uptake of ozone to aqueous

523

halogenide particles. Atmos. Chem. Phys. 2010, 10 (23), 11489-11500.

524

(19) Shrestha, M.; Luo, M.; Li, Y.; Xiang, B.; Xiong, W.; Grassian, V. H. Let there be light: stability

525

of palmitic acid monolayers at the air/salt water interface in the presence and absence of simulated

526

solar light and a photosensitizer. Chem. Sci. 2018, 9 (26), 5716-5723.

527

(20) Mochida, M.; Kitamori, Y.; Kawamura, K.; Nojiri, Y.; Suzuki, K. Fatty acids in the marine

528

atmosphere: Factors governing their concentrations and evaluation of organic films on sea-salt

529

particles. J. Geophys. Res.: Atmos. 2002, 107 (D17), 4325.

530

(21) Sicre, M. A.; Marty, J. C.; Saliot, A. n-Alkanes, fatty acid esters, and fatty acid salts in size

531

fractionated aerosols collected over the Mediterranean Sea. J. Geophys. Res.: Atmos. 1990, 95 (D4),

532

3649-3657.

533

(22) Cincinelli, A.; Stortini, A.; Perugini, M.; Checchini, L.; Lepri, L. Organic pollutants in

sea-534

surface microlayer and aerosol in the coastal environment of Leghorn—(Tyrrhenian Sea). Mar. Chem.

535

2001, 76 (1-2), 77-98.

536

(23) Ehrhardt, M.; Osterroht, C.; Petrick, G. Fatty-acid methyl esters dissolved in seawater and

537

associated with suspended particulate material. Mar. Chem. 1980, 10 (1), 67-76.

538

(24) Mukherjee, S.; Datta, A. Langmuir-Blodgett deposition selects carboxylate headgroup

539

coordination. Phys. Rev. E 2011, 84 (4), 041601.

540

(25) Tang, C. Y.; Huang, Z. S. A.; Allen, H. C. Binding of Mg2+ and Ca2+ to palmitic acid and

541

deprotonation of the COOH headgroup studied by vibrational sum frequency generation spectroscopy.

542

J. Phys. Chem. B 2010, 114 (51), 17068-17076. 543

(26) Gericke, A.; Brauner, J. W.; Erukulla, R. K.; Bittman, R.; Mendelsohn, R. In-situ investigation

544

of partially deuterated fatty acid and phospholipid monolayers at the air-water interface by IR

545

reflection-absorption spectroscopy. Thin Solid Films 1996, 284 (1-2), 428-431.

546

(27) Tang, C. Y.; Allen, H. C. Ionic binding of Na+ versus K+ to the carboxylic acid headgroup of

547

palmitic acid monolayers studied by vibrational sum frequency generation spectroscopy. J. Phys.

548

Chem. A 2009, 113 (26), 7383-7393. 549

(21)

(28) Teer, E.; Knobler, C. M.; Lautz, C.; Wurlitzer, S.; Kildae, J.; Fischer, T. M. Optical

550

measurements of the phase diagrams of Langmuir monolayers of fatty acid, ester, and alcohol

551

mixtures by brewster-angle microscopy. J. Chem. Phys. 1997, 106 (5), 1913-1920.

552

(29) Simon-Kutscher, J.; Gericke, A.; Huhnerfuss, H. Effect of bivalent Ba, Cu, Ni, and Zn cations

553

on the structure of octadecanoic acid monolayers at the air-water interface as determined by external

554

infrared reflection-absorption spectroscopy. Langmuir 1996, 12 (4), 1027-1034.

555

(30) Sinnamon, B. F.; Dluhy, R. A.; Barnes, G. T. Reflection-absorption FT-IR spectroscopy of

556

pentadecanoic acid at the air/water interface. Colloids Surf., A 1999, 146 (1-3), 49-61.

557

(31) Nikolova, G. S.; Zhang, L.; Chen, X.; Chi, L.; Haufe, G. Selective synthesis and self-organization

558

at the air/water interface of long chain fluorinated unsaturated ethyl esters and alcohols. Colloids Surf.,

559

A 2008, 317 (1-3), 414-420. 560

(32) Nieto-Suarez, M.; Vila-Romeu, N.; Prieto, I. Influence of different factors on the phase

561

transitions of non-ionic Langmuir monolayers. Appl. Surf. Sci. 2005, 246 (4), 387-391.

562

(33) Nieto-Suarez, M.; Vila-Romeu, N.; Dynarowicz-Latka, P.; Prieto, I. The influence of inorganic

563

ions on the properties of nonionic Langmuir monolayers. Colloids Surf., A 2004, 249 (1-3), 11-14.

564

(34) Pelletier, I.; Bourque, H.; Buffeteau, T.; Blaudez, D.; Desbat, B.; Pezolet, M. Study by infrared

565

spectroscopy of ultrathin films of behenic acid methyl ester on solid substrates and at the air/water

566

interface. J. Phys. Chem. B 2002, 106 (8), 1968-1976.

567

(35) Gericke, A.; Hühnerfussf, H. The conformational order and headgroup structure of long-chain

568

alkanoic acid ester monolayers at the air/water interface. Ber. Bunsenges. Phys. Chem. 1995, 99 (4),

569

641-650.

570

(36) Gericke, A.; Huhnerfuss, H. In-situ investigation of saturated long-chain fatty-acids at the

air-571

water-interface by external infared reflection-absorption spectrometry. J. Phys. Chem. 1993, 97 (49),

572

12899-12908.

573

(37) Brzozowska, A. M.; Duits, M. H. G.; Mugele, F. Stability of stearic acid monolayers on artificial

574

sea water. Colloids Surf., A 2012, 407, 38-48.

575

(38) Kester, D. R.; Duedall, I. W.; Connors, D. N.; Pytkowicz, R. M. Preparation of artificial seawater.

576

Limnol. Oceanogr. 1967, 12 (1), 176-179. 577

(39) Forestieri, S. D.; Staudt, S. M.; Kuborn, T. M.; Faber, K.; Ruehl, C. R.; Bertram, T. H.; Cappa,

578

C. D. Establishing the impact of model surfactants on cloud condensation nuclei activity of sea spray

579

aerosol mimics. Atmos. Chem. Phys. 2018, 18 (15), 10985-11005.

580

(40) Khattari, Z.; Sayyed, M. I.; Qashou, S. I.; Fasfous, I.; Al-Abdullah, T.; Maghrabi, M. Interfacial

581

behavior of myristic acid in mixtures with DMPC and cholesterol. Chem. Phys. 2017, 490, 106-114.

582

(41) Hao, C. C.; Sun, R. G.; Zhang, J. Mixed monolayers of DOPC and palmitic acid at the liquid-air

583

interface. Colloids Surf., B 2013, 112, 441-445.

584

(42) Griffith, E. C.; Guizado, T. R.; Pimentel, A. S.; Tyndall, G. S.; Vaida, V. Oxidized

aromatic-585

aliphatic mixed films at the air-aqueous solution interface. J. Phys. Chem. C 2013, 117 (43),

22341-586

22350.

587

(43) Minh Dinh, P.; Lee, J.; Shin, K. Collapsed States of Langmuir Monolayers. J. Oleo Sci. 2016,

588

65 (5), 385-397. 589

(44) Kaganer, V. M.; Mohwald, H.; Dutta, P. Structure and phase transitions in Langmuir monolayers.

590

Rev. Mod. Phys. 1999, 71 (3), 779-819. 591

(45) Hasegawa, T.; Nishijo, J.; Watanabe, M.; Umemura, J.; Ma, Y. Q.; Sui, G. D.; Huo, Q.; Leblanc,

592

R. M. Characteristics of long-chain fatty acid monolayers studied by infrared external-reflection

593

spectroscopy. Langmuir 2002, 18 (12), 4758-4764.

594

(46) Adams, E. M.; Allen, H. C. Palmitic acid on salt subphases and in mixed monolayers of

595

cerebrosides: Application to atmospheric aerosol chemistry. Atmosphere 2013, 4 (4), 315-336.

(22)

(47) Nutting, G. C.; Harkins, W. D. Pressure-area relations of fatty acid and alcohol monolayers. J.

597

Am. Chem. Soc. 1939, 61, 1180-1187. 598

(48) Sierra-Hernandez, M. R.; Allen, H. C. Incorporation and exclusion of long chain alkyl halides in

599

fatty acid monolayers at the air-water interface. Langmuir 2010, 26 (24), 18806-18816.

600

(49) Brzozowska, A.; Mugele, F.; Duits, M. Stability and interactions in mixed monolayers of fatty

601

acid derivatives on artificial sea water. Colloids Surf., A 2013, 433, 200-211.

602

(50) Yue, X. L.; Steffen, P.; Dobner, B.; Brezesinski, G.; Mohwald, H. Monolayers of mono- and

603

bipolar palmitic acid derivatives. Colloids Surf., A 2004, 250 (1-3), 57-65.

604

(51) Gershevitz, O.; Sukenik, C. N. In situ FTIR-ATR analysis and titration of carboxylic

acid-605

terminated SAMs. J. Am. Chem. Soc. 2004, 126 (2), 482-483.

606

(52) Johann, R.; Vollhardt, D. Texture features of long-chain fatty acid monolayers at high pH of the

607

aqueous subphase. Mater. Sci. Eng., C 1999, 8, 35-42.

608

(53) Wang, L.; Jacobi, S.; Sun, J.; Overs, M.; Fuchs, H.; Schaefer, H. J.; Zhang, X.; Shen, J.; Chi, L.

609

Anisotropic aggregation and phase transition in Langmuir monolayers of methyl/ethyl esters of 2,

3-610

dihydroxy fatty acids. J. Colloid Interface Sci. 2005, 285 (2), 814-820.

611

(54) Wang, Y.; Du, X.; Guo, L.; Liu, H. Chain orientation and headgroup structure in Langmuir

612

monolayers of stearic acid and metal stearate (Ag, Co, Zn, and Pb) studied by infrared

reflection-613

absorption spectroscopy. J. Chem. Phys. 2006, 124 (13), 134706.

614

(55) Chen, Q. B.; Kang, X. L.; Li, R.; Du, X. Z.; Shang, Y. Z.; Liu, H. L.; Hu, Y. Structure of the

615

Complex Monolayer of Gemini Surfactant and DNA at the Air/Water Interface. Langmuir 2012, 28

616

(7), 3429-3438.

617

(56) Pelletier, I.; Laurin, I.; Buffeteau, T.; Desbat, B.; Pézolet, M. Infrared study of the molecular

618

orientation in ultrathin films of behenic acid methyl ester: comparison between single

Langmuir-619

Blodgett monolayers and spin-coated multilayers. Langmuir 2003, 19 (4), 1189-1195.

620

(57) Muro, M.; Itoh, Y.; Hasegawa, T. A conformation and orientation model of the carboxylic group

621

of fatty acids dependent on chain length in a Langmuir monolayer film studied by

polarization-622

modulation infrared reflection absorption spectroscopy. J. Phys. Chem. B 2010, 114 (35),

11496-623

11501.

624

(58) Levin, I. W.; Thompson, T. E.; Barenholz, Y.; Huang, C. Two types of hydrocarbon chain

625

interdigitation in sphingomyelin bilayers. Biochemistry 1985, 24 (22), 6282-6286.

626

(59) Aoki, P. H. B.; Morato, L. F. C.; Pavinatto, F. J.; Nobre, T. M.; Constantino, C. J. L.; Oliveira,

627

O. N., Jr. Molecular-Level Modifications Induced by Photo-Oxidation of Lipid Monolayers

628

Interacting with Erythrosin. Langmuir 2016, 32 (15), 3766-3773.

629

(60) Huang, C. H.; Lapides, J. R.; Levin, I. W. Phase-transition behavior of saturated, symmetric

630

chain phospholipid-bilayer dispersions determined by Raman-spectroscopy-correlation between

631

spectral and thermodynamic parameters. J. Am. Chem. Soc. 1982, 104 (22), 5926-5930.

632

(61) Sakai, H.; Umemura, J. Molecular orientation in Langmuir films of 12-hydroxystearic acid

633

studied by infrared external-reflection spectroscopy. Langmuir 1998, 14 (21), 6249-6255.

634

(62) Le Calvez, E.; Blaudez, D.; Buffeteau, T.; Desbat, B. Effect of cations on the dissociation of

635

arachidic acid monolayers on water studied by polarization-modulated infrared reflection-absorption

636

spectroscopy. Langmuir 2001, 17 (3), 670-674.

637

(63) Avila, L.; Saraiva, S.; Oliveira, J. Stability and collapse of monolayers of stearic acid and the

638

effect of electrolytes in the subphase. Colloids Surf., A 1999, 154 (1-2), 209-217.

639

(64) Griffith, E. C.; Adams, E. M.; Allen, H. C.; Vaida, V. Hydrophobic collapse of a stearic acid

640

film by adsorbed l-phenylalanine at the air-water interface. J. Phys. Chem. B 2012, 116 (27),

7849-641

7857.

642

(65) Gilman, J. B.; Tervahattu, H.; Vaida, V. Interfacial properties of mixed films of long-chain

643

organics at the air-water interface. Atmos. Environ. 2006, 40 (34), 6606-6614.

(23)

(66) Gilman, J. B.; Eliason, T. L.; Fast, A.; Vaida, V. Selectivity and stability of organic films at the

645

air-aqueous interface. J. Colloid Interface Sci. 2004, 280 (1), 234-243.

646

(67) Tervahattu, H.; Juhanoja, J.; Kupiainen, K. Identification of an organic coating on marine aerosol

647

particles by TOF-SIMS. J. Geophys. Res.: Atmos. 2002, 107 (D16), 4319.

648

(68) Noziere, B.; Baduel, C.; Jaffrezo, J.-L. The dynamic surface tension of atmospheric aerosol

649

surfactants reveals new aspects of cloud activation. Nat. Commun. 2014, 5, 3335.

650

(69) Farmer, D. K.; Cappa, C. D.; Kreidenweis, S. M. Atmospheric Processes and Their Controlling

651

Influence on Cloud Condensation Nuclei Activity. Chem. Rev. 2015, 115 (10), 4199-4217.

652

(70) Gorbunov, B.; Hamilton, R.; Clegg, N.; Toumi, R. Water nucleation on aerosol particles

653

containing both organic and soluble inorganic substances. Atmos. Res. 1998, 48, 271-283.

654 655

Figure

Table 1. Composition of the artificial seawater  115
Figure 1. These  π− A isotherms show remarkable changes along with alkyl chain-length, head-groups, 158
Figure 2. IRRAS spectra (2820-2950 cm -1 ) of the fatty acid ((a), (b)) and fatty acid methyl ester ((c),  280
Table 2. The peak-height intensity ratio between the antisymmetric and symmetric bands of the CH 2

Références

Documents relatifs

Recently a few patients have been described with mild MMAuria associated with a mutation of the MMCoA epimerase gene (MCEE), but these patients had no consistent clinical phenotype,

Figure 2: Relative invasion (average +/- SD) in IPI-2I cells of Salmonella Typhimurium grown in LB broth, supplemented with sub-MIC concentrations of different fatty

of dietary fibers, namely the short chain fatty acids (SCFAs) linking host nutrition to intestinal 26... SCFAs are an important fuel for intestinal epithelial cells (IEC) and

An appropriate offline DP-SEO algorithm permitting to simultaneously generate an optimal speed profile and its power split strategy is proposed in this paper in order to ensure:

En fonction de la granularité atteinte par les dépôts les plus grossiers, cette séquence-type se décline en trois variantes (fine, moyenne et grossière) qui, à elles

It has been indeed reported that oral ingestion of TG in rats decreased food intake through hepatic fatty acid oxidation CD36 Mediates Fatty Acids Effects on Food Intake... It has

Moreover, in vitro assessment of the antibacterial activity of marbofloxacin against both bacterial species demonstrated that the inoculum effect, that corresponds to the decrease

م‌ - لوادجلا ةمئاق : مقر لودجلا لودجلا ناونع ةحفصلا 1 ثيدحلا جاينملاك ـيدقلا جمانربلا فيب ةنراقم حضكي ةيضايرلاك ةيندبلا ةيبرتل 20